Electronic Structure and Optical Properties of an Alternated Fluorene

Mar 8, 2012 - Theoretical calculations (density functional theory and semiempirical methodologies) used to simulate the geometry of some oligomers and...
0 downloads 10 Views 2MB Size
Article pubs.acs.org/JPCA

Electronic Structure and Optical Properties of an Alternated Fluorene−Benzothiadiazole Copolymer: Interplay between Experimental and Theoretical Data Paula C. Rodrigues,†,‡ Leonardo S. Berlim,§,‡ Diego Azevedo,#,‡ Nestor C. Saavedra,∥ Paras N. Prasad,⊥ Wido H. Schreiner,§ Teresa D. Z. Atvars,# and Leni Akcelrud*,† †

Paulo Scarpa Polymer Laboratory, Federal University of Parana (LaPPS), CP 19081, CEP 81531-980, Curitiba, Paraná, Brazil Laboratory of Surfaces and Interfaces, Federal University of Parana (LSI), CP 19081, CEP 81531-980, Curitiba, Paraná, Brazil ∥ Physics Department, Federal Technological University of Parana, Av. Sete de Setembro, 3165, CEP 80230-901, Curitiba, Paraná, Brazil ⊥ Institute for Lasers, Photonics and Biophotonics (ILPB)−University at Buffalo, State University of New York, 458 Natural Sciences Complex-North Campus, Buffalo, New York 14260, United States # Institute of Chemistry, State University of Campinas (UNICAMP), P.O. Box 6154, Campinas, SP, CEP 13084-971, Brazil §

S Supporting Information *

ABSTRACT: The donor−acceptor copolymer containing benzothiadiazole (electron acceptor), linked to functionalized fluorene (electron donor), [poly[9,9-bis(3′-(tert-butyl propanoate))fluorene-co-4,7-(2,1,3-benzothiadiazole)] (LaPPS40), was synthesized through the Suzuki route. The polymer was characterized by scanning electron microscopy, gel permeation chromatography, NMR, thermal analysis, cyclic voltammetry, X-ray photoelectron spectroscopy, UV−vis spectrometry, and photophysical measurements. Theoretical calculations (density functional theory and semiempirical methodologies) used to simulate the geometry of some oligomers and the dipole moments of molecular orbitals involved were in excellent agreement with experimental results. Using such data, the higher energy absorption band was attributed to the π−π* (S0 → S4) transition of the fluorene units and the lower lying band was attributed to the intramolecular (ICT) (S0 → S1) charge transfer between acceptor (benzothiadiazole) and donor groups (fluorene) (D−A structure). The ICT character of this band was confirmed by its solvatochromic properties using solvents with different dielectric properties, and this behavior could be well described by the Lippert−Mataga equation. To explain the solvatochromic behavior, both the magnitude and orientation of the dipole moments in the electronic ground state and in the excited state were analyzed using the theoretical data. According to these data, the change in magnitude of the dipole moments was very small for both transitions but the spatial orientation changed remarkably for the lower energy band ascribed to the ICT band.

1. INTRODUCTION In the steadily growing field of polymer solar cells several types of polymer systems and device architectures have emerged as attractive means to improve the performance and chemical stability of the devices.1 There are two main approaches to achieve this goal: one is the blending of polymer materials to obtain new properties or to combine the characteristics of the parent components, producing devices with enhanced performance. The other involves developing new materials, in general copolymers, with better properties for specific applications.1−3 In this last category, molecular engineering of conjugated polymers should be performed aiming the reduction of the band gap and at the same time modulating HOMO and LUMO energy levels to optimal values, both fundamental requirements for solar cells applications.4 Under this perspective, it is practically impossible to achieve these properties using a polymer with only one repeating unit, and copolymers represent an alternative in this line of pursuit. One particular © 2012 American Chemical Society

type of promising structure that has been explored with success is characterized by the presence of groups with different electron affinities forming a type of donor−acceptor (D−A) system which may favor charge separation processes.5 Since at least two comonomers must be used in this type of copolymers, their electronic absorption bands may cover a wider range of the visible spectrum, enhancing the ability to absorb solar light, making them better light collectors than polymers with only one chromophore. The interaction between an electron acceptor with a donor6 facilitates photoinduced charge separation in photovoltaic devices. Moreover, the D−A structure allows the desired tuning of the energy levels, with systematic variation in the polymer electronic structure. The origin of the well-separated dual-absorption band sometimes Received: November 16, 2011 Revised: February 27, 2012 Published: March 8, 2012 3681

dx.doi.org/10.1021/jp211042v | J. Phys. Chem. A 2012, 116, 3681−3690

The Journal of Physical Chemistry A

Article

Figure 1. Chemical route for synthesis of poly[9,9-bis(3′-(tert-butyl propanoate))fluorene-co-4,7-(2,1,3-benzothiadiazole)] (LaPPS40).

encountered in semiconducting polymers using the “donor− acceptor” concept has been reviewed,5 and the present contribution aims to add further insight to this important subject by describing the correlations found between structure and optical properties of the D−A copolymer, taking into account theoretical and experimental data. The D−A character of a chemical structure can be studied by theoretical and experimental methodologies. From the experimental point of view, one possible methodology is the photoluminescence technique, where the D−A molecule can depict solvatochromic properties in solvents with different polarities. The larger the changes of the dipole moments by photoexcitation, the higher the spectral changes will be either in terms of band profile or in terms of peak positions. For several systems these solvatochromic properties can be well represented by the Lippert−Mataga equation,7,8 which associates the spectral changes with modification of the dipole moments of both the electronic ground state and the electronic excited state. Although these changes can be experimentally determined, the individual dipole moments of each component or their spatial orientation cannot be obtained with this data set. There are several methodologies for theoretical studies of chemical compounds attempting to describe their optical properties.9,10 In general, quantum mechanical calculations may be used for complex and larger systems with the use of adequate methodology, taking into account that often the optical properties of polymers can be simulated using oligomers. Here the theoretical study was performed by computational quantum mechanical calculations initially using the density functional theory (DFT) method11 to optimize the geometry of a single repeating unit. Having the optimized geometry of this structure, using semiempirical quantum mechanical calculations, the geometry of oligomers was also optimized. The extent of the D−A character was probed by analyzing the changes of dipole moments in both the electronic ground and the excited states, which were further correlated with the experimental data of optical properties in different solvents. To attain a significant increase in the intramolecular charge transfer band, adequate combination of donor and acceptor groups is crucial.12 Examples of electron-accepting species include quinoline, 2,1,3-benzothiadiazole, thiazole, 1,3,4-

thiadiazole, pyridazine, and 1,3,5-triazine as electron-accepting species,3,13−15 whereas heteroaromatic rings such as thiophene and pyrrole, along with other units like fluorene,16 carbazole,17 4H-cyclopenta[2,1-b:3,4-b′]dithiophene (CPDT), and dithieno[3,2-b:2′,3′-d]silole (DTS)18 polymers have been explored as donor species. The D−A strategy has afforded results in PCEs reaching the 4−7% range.19−21 In this work two classical units in D−A copolymer architecture were used to get insights related to the changes of dipole moments by photoexcitation and for describing the D−A character of the electronic transitions: the fluorene moiety as donor and benzothiadiazole as acceptor units, composing the alternated poly[9,9-bis(3′-(tert-butyl propanoate))fluorene-co-4,7-(2,1,3-benzothiadiazole)] (LaPPS40). Insertion of functional groups in the side chains of polyfluorene aimed to facilitate processing and expand applications of the material. The (tert-butyl propanoate) branches can easily be converted to carboxyl groups, rendering the polymer especially suitable for interaction with the titanium oxide in dye-type solar cells. This kind of modification has been addressed in similar cases, as, for example, fabrication of anionic polyfluorene focusing on its use in optoelectronic devices as LEDs22 as well as in biochemical sensors.23,24 The changes of the optical properties were studied using both the experimental approach determining some photophysical properties of the copolymer and theoretical methodologies to support the experimental data.

2. EXPERIMENTAL SECTION Materials. The chemicals used were all purchased from Aldrich and used as received, without further purification, unless when described in the specific chemical procedure. Solvents (chloroform, toluene, N-methyl pyrrolidone (NMP), and o-dichlorobenzene (o-DCB)) were from Aldrich, HPLC grade and used as received. Tetrahydrofurane (THF) (Aldrich, HPLC) was purified by distillation. Chemical Procedures. Poly [9,9-bis (3 ′ -(t e r tbutylpropanoate))fluorene-co-4,7-(2,1,3-benzothiadiazole)] (LaPPS40) was synthesized by the Suzuki route (Figure 1) as described elsewhere.25 2,7-Dibromo-9,9-bis(3-(tert-butyl propanoate))fluorene (1). To a two-necked round-bottom flask, 2,7-dibromofluorene 3682

dx.doi.org/10.1021/jp211042v | J. Phys. Chem. A 2012, 116, 3681−3690

The Journal of Physical Chemistry A

Article

Elemental analysis (Eager 200 model, C, H, N, S) was carried out to determine the composition of the precursors and LaPPS40 copolymer. The thermal decomposition temperature of the prepared polymer was determined by a Netzsch Analyzer TG 209. For the thermogravimetric tests (TGA) a heating rate of 20 °C/min was used with a nitrogen flow of 15 mL/min from room temperature to 450 °C. The DSC equipment used was a Netzsch DSC 204 F1. All samples were heated from 20 to 250 °C at a rate of 10 °C/min in nitrogen atmosphere and then cooled to 20 at 10 °C/min. This procedure was repeated, and the second run was recorded. UV−vis spectra were taken in a Shimadzu spectrophotometer model NIR 3101. Steady-state fluorescence spectroscopy was performed in Shimadzu spectrophotometer model RF5301-PC, using a square cuvette of 1 cm for solutions. The spectral range was from 250 to 500 nm for the excitation spectra and from 300 to 650 nm for emission spectra (λexc = 445 nm). Time-resolved emission was performed using a F900 (Edinburg Analytical Instruments) instrument using a singlephoton counting technique, using pulsed diode lasers (λexc =375 and 470 nm) and a MCP-PMT as a detector. Decays we recorded in the emission spectral range from 500 to 600 nm. Data treatment was done using exponential functions by the software provided by Edinburgh Instruments

(3.03 g, 12.4 mmol), tetrabutylammonium bromide (105 mg, 0.33 mmol), and 20 mL of toluene were added under Ar atmosphere. After 15 min, a solution of 50 wt % aqueous NaOH was added dropwise and the solution was stirred for 20 min. After that, tert-butyl acrylate (6.56 g, 51.2 mmol) was added dropwise and the mixture was stirred at 25 °C overnight. After this time, the reaction product was extracted with dichloromethane, washed three times with water and saturated NaCl aqueous solution, and dried with anhydrous MgSO4. After solvent removal the product was purified in a chromatographic column to give a white solid (1) with a 75% yield. 1 H NMR (CDCl3, δ): 7.54−7.48 (m, 6H); 2.30 (t, 4H), 1.47 (t, 4H), 1.30 (s, 18H). 13C NMR (CDCl3, δ): 172.22; 149.96; 139.09; 131.06; 126.48; 122.07; 121.45; 80.48; 54.07; 34.41; 29.85; 28.02. Anal. Calcd for C27H32Br2O4: C, 55.88; H, 5.56. Found: C, 55.67; H, 5.45. 2-(4,4,5,5-Tetramethyl-1,3,2-dioxaborolan-2-yl)-9,9-bis(3(tert-butyl propanoate))fluorene (2). A mixture containing 3.27 g (6.5 mmol) of 1, 2.7 g (27.5 mmol) of KOAc, and 4.2 g (16.5 mmol) of bis(pinacolato)diboron in 40 mL of dry DMF was placed in a flask. After stirring for 20 min under Ar atmosphere, 150 mg of [1,1′-bis(diphenylphosphine)ferrocene] dichloropalladium(II) was added quickly. The mixture was stirred overnight at a temperature of 90 °C. After this period, the mixture was cooled, placed in water, and extracted with dichloromethane. The organic phase was washed with water and dried with MgSO4. After removing the solvent, the residue was purified with column chromatography, yielding compound 3 as a white solid (yield 57%). 1 H NMR (CDCl3, δ): 7.80−7.72 (m, 6H); 2.38 (t, 4H), 1.42−1.39 (m, 28H), 1.30 (s, 18H). 13C NMR (CDCl3, δ): 172.88; 147.89; 143.83; 134.36; 129.03; 128.40; 119.69; 83.92; 79.97; 53.54; 34.48; 29.99; 28.04; 24.96. Anal. Calcd for C39H56B2O8: C, 69.45; H, 8.37. Found: C, 67.78; H, 8.78. Poly[9,9-bis(3′-(tert-butyl propanoate))fluorene-co-4,7(2,1,3-benzothiadiazole)]−LaPPS40. To a 25 mL flask 210 mg (0.31 mmol) of 4,7-dibromo-2,1,3-benzothiadiazole, 91 mg (0.31 mmol) of 2, 8 mg of Pd (Ph3P)4, and 515 mg (3.7 mmol) of potassium carbonate were added. Then a mixture of water and toluene was added to the flask and Ar was bubbled into the system for 20 min. The mixture was heated to 90 °C for 24 h under Ar and after this period precipitated in methanol. Then the polymer was filtered and washed with methanol and acetone. Purification was performed by Sohxlet extraction, yielding a fibrous yellow solid (yield 57%). 1 H NMR (CDCl3, δ): 8.18−7.98 (br, 8H); 2.55 (br, 4H); 1.79 (m, 4H); 1.33 (s, 18H). 13C NMR (CDCl3, δ): 172.90; 154.25; 149.22; 140.96; 137.04; 133.35; 128.22; 129.27; 124.04; 120.38; 80.18; 54.03; 34.68; 30.32; 28.08. IR:ν(̅ CO)ester 1728 cm −1 ; ν̅ (C−O)ester 1149 cm −1 . Anal. Calcd for C33H36N2O4S: C, 71.19; H, 6.52; N, 5.03; S, 5.76. Found: C, 67.98; H, 5.92; N, 5.48; S, 4.26. Equipment. The molar mass measurements were performed with a Waters 2690 gel permeation chromatograph, equipped with a Waters 996 refraction index photodiode detector. The calibration curve was built with polystyrene standards and tetrahydrofurane (THF) as eluant at a flow rate of 1.0 mL/min at 30 °C. 1 H and 13C NMR spectra were recorded in chloroform on a Bruker system operating at 400 MHz at room temperature. Samples for FTIR measurements were cast onto polished silicon disks. Spectra were run in a Nicolet Magna-TR 560 infrared spectrophotometer.

N

F (t ) =

⎛ −t ⎞ ⎟ ⎝ τi ⎠

∑ Bi exp⎜ i=1

(1)

where Bi is a pre-exponential factor representing the fractional contribution to the time-resolved decay of the component with a lifetime τi and t is the time. Samples were saturated with N2. Ludox was used as scatterer to register the instrumental response. Deconvolution of the laser pulse was performed, optimizing the random residual distribution to achieve a χ2 parameter close to unity.26,27 Decays were initially recorded using pulsed diode lasers for excitation (375 and 470 nm) and collected with emission at 550 nm. Experiments of timeresolved emission spectroscopy were performed using the same excitations and tuning the emission from 500 to 700 nm in intervals of 10 nm. A potentiostat/galvanostat PAR 273A was used for electrochemical characterization with a three-electrode cell: working electrode (Pt), reference electrode (Ag/Ag+ in acetonitrile), counter electrode (Pt), supporting electrolyte (0.1 mol L−1 tetrabutylammonium hexafluorophosphate in acetonitrile), and scanning rate 50 mV/s. The ferrocenium/ferrocene pair was used as internal standard. Scanning electron microscopy was done with a Phenom FEI desktop instrument operating at 5 kV. The LAPPS40 film sample was coated with a sputtered gold film, and images were obtained by backscattered electrons. XPS spectra were collected on a VG Microtech ESCA3000 system using Al Kα X-rays. The system has a 250 mm semihemispherical analyzer with 9 channeltron detectors. The ultimate vacuum of the machine is 3 × 10−10 mbar, and the overall resolution is 0.8 eV. Survey and high-resolution spectra were obtained with, respectively, 50 and 20 eV bandpass energies. Quantification of the elements on the sample surface was done using the proprietary system software. Deconvolution of the high-resolution spectra was done using XPS International software.28 The Shirley method was used for background 3683

dx.doi.org/10.1021/jp211042v | J. Phys. Chem. A 2012, 116, 3681−3690

The Journal of Physical Chemistry A

Article

presenting a peculiar micrometric ribbon-like structure. The ribbons are composed of nanometric structures with porous characteristics. The amorphous nature of the material was confirmed by X-ray diffraction analysis (diffractogram not shown). The chemical structure was confirmed by 1H NMR, 13C NMR, and IR spectroscopies, as presented in the Experimental Section and shown in the Supporting Information. The thermal stability of the polymer was studied using TGA, showing that the LaPPS40 (ester form) is thermally stable until 200 °C under nitrogen atmosphere. At this temperature, a mass loss about 20% was observed, assigned to breaking of the ester link (Table 1). No other thermal transitions were observed in DSC

subtraction of the spectra. The C1s line for C−C/C−H bonds at 285 eV29 was used for all spectra binding energy corrections. Theoretical Methods. Electronic and geometric properties of some oligomers with fluorene and benzothiadiazole groups were studied using ab initio and semiempirical quantum mechanical calculations. First, the geometry of the fluorene− benzodiathiazole unit was optimized with the DFT method using the Becke three-parameter, Lee−Yang−Parr functional (B3LYP),30 which is a hybrid functional of exact (Hartree− Fock) exchange with local and gradient-corrected exchanges and correlation terms used in calculations of very large molecules.31 A set of Gaussian Split-based Valence wave function basis was used.32 For purposes of reliability in the calculations the geometry was previously optimized using the basic Gaussian split-valence 6-31G* basis, which gives good ground state geometries for conjugated polymers.17,33 These simulations were performed using the GAUSSIAN 0930 package. DFT/B3LYP calculations with the 6-31G* basis set running GAUSSIAN 09 were performed. After that the geometry was simulated for oligomers with 1− 6mers using semiempirical calculations. Each oligomer represented the previous optimized geometry of a smaller one increased by adding a new unit at both sides of the molecule. Due to lower computational load, all computational results discussed in this calculation were obtained performing calculations with the PM6 Hamiltonian contained in the MOPAC package.34 Finally, simulation of the LaPPS40 optical absorption spectrum was performed using ZINDO (Zerner’s intermediate neglect of differential overlap/single)30 running under the GAUSSIAN 09 program. The ZINDO transition energies were weighted by the oscillator strength values. The theoretically calculated values were used to assign the orbital related with the experimental data of the HOMO−LUMO optical transition

Table 1. Results of XPS Analysis for the LaPPS40 Polymer and Some Theoretically Data for an Oligomer with Three Repeating Units

a

element

XPS (%)

theoretical (%)

binding energy (eV) (XPS)

binding energy (eV) (theoretical)a

S C N O C−C/C−H C−O CO

2.5 81.2 4.4 11.9 70.4 5.8 5.0

2.5 82.5 5 10 72.5 5.0 5.0

165.37/166.54 285−289 399.75 533.0 285.0 287.0 289.0

162.85/163.24 276.66−280.82 391.38 521.63 276.66 277.79 280.82

Estimated using the ab initio method.

measurements in the 50−300 °C temperature range (shown in the Supporting Information). LaPPS40 Electronic Properties. Table 1 shows some results of the XPS measurements. The Sp3/2 and Sp1/2, N1s, and O1s binding energies were in excellent agreement with published results for similarly bonded polymers.35 The binding energies of the various carbon bonds also agree with the relative shifts to the C−C/C−H bonds at 285 eV. Figure 3 shows the high-resolution spectrum for the C1s line of the LaPPS40 polymer. Deconvolution of the spectrum leads to quantification of the C−C/C−H, C−O, and CO bonds. The calculated

3. RESULTS AND DISCUSSION Polymer Characterization. The molar mass of poly[9,9bis(3′-(tert-butyl propanoate))fluorene-co-4,7-(2,1,3-benzothiadiazole)] (LaPPS40) as determined by GPC using polystyrene as standards was M̅ w = 32 000 g mol−1 with a polydispersity (PDI) of 1.40. The polymer showed good solubility in organic solvents such as CHCl3, THF, and dichloromethane, among others. It could be collected as fibrous strands and formed homogeneous, self-supporting, and continuous films by casting from solutions. These properties indicated that the real molar mass of the polymer was probably larger than that obtained from GPC data, based on polystyrene standards. Figure 2 shows typical SEM images of the solid state polymer,

Figure 3. High-resolution XPS spectrum of the C1s line of the LaPPS40 polymer: (A) C−C/C−H bonds, (B) C−O bonds, and (C) CO bonds.

Figure 2. SEM images of the LaPPS40 in as-cast solid state. 3684

dx.doi.org/10.1021/jp211042v | J. Phys. Chem. A 2012, 116, 3681−3690

The Journal of Physical Chemistry A

Article

atomic percentage and binding energies of these bonds are shown in Table 1. Figure 4 shows the valence band spectrum of the LaPPS40 polymer obtained with XPS, adequately shifted in energy. The

Figure 4. Valence band spectrum of the LaPPS40 polymer obtained by XPS compared to the density of states of the theoretical simulation.

valence band below the Fermi level is dominated by the C2p and C2s bands, and the main peaks of the spectrum are reproduced by the theoretical density of states calculation shown in the same figure. The overall agreement of quantification for the XPS results as compared to the theoretically expected was very good. Since the DFT calculation and semiempirical method showed similar results for optimization of the repeating unit and due to lower computational load the computational results discussed were obtained performing calculations with the PM6 Hamiltonian (semiempirical method). The geometry of the oligomer containing three and six repeating units is shown in Figure 5a and 5e, respectively. The optimized geometry and corresponding energies of the HOMO−LUMO orbitals are shown on Figure 5b−d (for three units) and 5f−h (for six units). The calculation indicates that the fluorene units in the nmers are in a planar configuration, but their relative orientation relative to the benzothiadiazole groups is almost orthogonally oriented in the electronic ground state (Figure 5a and 5e). Since the simulations using three or six repeating units show the same geometry and similar electronic density, due to computational resources the compound model with three repeating units was used as the LaPPS40 representative unit for absorption transitions (Figure 10) and core energy levels (Figure 4). The moduli of the dipole moments for every state did not vary significantly, as shown in Figure 5 (p values), but their spatial orientation suffered appreciable changes as shown at Table 2 by analysis of the x,y,z components. Single-point calculations were initially performed, and the values for the HOMO and LUMO were obtained for oligomers with 1−6mers using the same functional and the same wave function basis set. The convergence of the energy values is shown in Figure 6a. It is noteworthy that the theoretical data were also consistent with those determined experimentally by XPS for the valence band spectrum of LaPPS40 (Table 1). Figure 6b shows the theoretical energy band gap as a function of the inverse of the number of repeating units (1/n). The relationship between these two parameters can be described by an exponential approach given by the Meyertype equation.36 It indicates that the polymer reaches a better convergence for the hexamer, and that value could be extrapolated to the “ideal”infinite polymer (see Supporting Information). Nevertheless, the three-unit model was used

Figure 5. Electronic ground state optimized geometry for an oligomer with three (a−d) and six repeating units (e−h). Charge density distribution for the HOMO orbital (b and f), S1 (LUMO) orbital (c and g), and S4 (LUMO+3) orbital (d and h) (isovalues 0.01). (Inset) Arrows representing the Cartesian dipole moment orientation for 3 repeating units (p).

since it saves computational resources and provided good agreement with experimental data. The energy of the HOMO and LUMO levels for the three repeating units is satisfactory when compared with the experimental data from cyclic voltammetry and from the optical absorption spectrum (see below). This fact does not mean that the effective conjugation of the polymer is reached with only three monomeric units, but it gives support to perform the quantum mechanical calculations over three repeating units. On the basis of the semiempirical optimized geometry for the oligomer with two, three, and four repeating units and on theoretical calculations for the orbital energies, evolution of the theoretical absorbance spectra is shown in Figure 7. The spectral shape with a “camel back” character is known to be a signature of the donor−acceptor copolymer structure with an ICT state.10 Using the calculated data for the oligomer with three repeating units the contribution of the valence molecular orbitals to the lower energy electronic transitions was analyzed. Assuming that the transition is purely electronic in vacuum and that no vibrational coupling is involved, the higher energy absorption band may be attributed to the π−π* transition (S0 → S4; with a major contribution of HOMO → LUMO+3), with an oscillator strength of 1.0604 (Table 3). This transition brings about an excited state with a calculated electric dipole of 3685

dx.doi.org/10.1021/jp211042v | J. Phys. Chem. A 2012, 116, 3681−3690

The Journal of Physical Chemistry A

Article

Table 2. Dipole Moments for the Excited and Ground State of LaPPS40 Repeating Units dipole moment components (Debye) pX excited state ground state S0

S1 S4

pY

pZ

dipole moment p (Debye)

n=3

n=6

n=3

n=6

n=3

n=6

n=3

n=6

4.148 −3.374 −3.531

7.618 −1.474 6.941

−0.018 0.028 1.184

0.052 −0.012 0.320

0.032 −0.023 −1.332

0.188 −0.175 0.851

4.14 3.37 3.95

7.62 1.48 7.00

Table 3. Major Transition Contributions to Theoretically Simulated Electronic Absorption Spectra for an Oligomer with Three Repeating Units theoretical wavelength (nm)

oscillator strength

326

1.0604

461

1.1319

transition

contribution (%)

HOMO → LUMO+3 HOMO−2 → LUMO+5 HOMO−1 → LUMO+4 HOMO → LUMO+4 HOMO−1 → LUMO+2 HOMO → LUMO HOMO−3 → LUMO+2 HOMO → LUMO+2

17 6 7 9 13 38 6 7

for this excited state clearly shows a charge transfer from the fluorene to the benzothiadiazole units, and thus, the transition undergoes a clear character of an intramolecular charge transfer (ICT) between the donor and the acceptor. It is noteworthy that although the magnitude of the calculated dipole moments does not change much (from 3.95 to 4.14 D) there is an inversion of their spatial orientation. LaPPS40 Optical Properties. Figure 8 shows the optical electronic absorption spectra of LaPPS40 in chloroform Figure 6. Calculated HOMO and LUMO energy levels for LaPPS40 repeating units (a), and theoretical band energy gap versus 1/n (where n is the number of repeating units) (b).

Figure 8. Electronic absorption spectrum of LaPPS40 solution in chloroform (10−5 mol L−1).

solution (10−5 mol L−1). The spectrum displays two absorption maxima located at λmax = 317 and 445 nm (3.91 and 2.79 eV, respectively), relatively close to the calculated values (Table 2). The oxidation and reduction potentials of LaPPS40 were also determined by cyclic voltammetry (Supporting Information), and the values for the HOMO and LUMO levels were determined as −5.60 and −3.07 eV, respectively, using the equations1,37,38

Figure 7. Theoretical absorption spectra of LaPPS40 repeating units (n).

3.37 D, lower than that of the electronic ground state 3.95 D, but the dipole moments maintained a similar spatial orientation. The charge distribution map for this excited state can be seen at Figure 5d, where the electronic cloud is spread over the main chain in a similar fashion as in the electronic ground state (Figure 5b). The lower energy absorption band may be ascribed to the transition between S0 and S1, with a major contribution of HOMO→LUMO (Figure 5c). Comparing with the HOMO (Figure 5b), the charge distribution map 3686

Ea(ELUMO) = (E′red + 4.5) eV

(2)

Ip(EHOMO) = (E′ox + 4.5) eV

(3)

dx.doi.org/10.1021/jp211042v | J. Phys. Chem. A 2012, 116, 3681−3690

The Journal of Physical Chemistry A

Article

where Ea is the electron affinity and Ip is the ionization potential. E′red and E′ox are the onset reduction and oxidation potentials versus the SHE reference electrode. Taking into account the approximations used in the theoretical information, these two bands can be ascribed to transitions from HOMO → LUMO+3 and to the HOMO → LUMO where a charge transfer occurs from the fluorene to the benzothiadiazole units. To confirm experimentally the ICT character of this polymer, the electronic absorption spectra of LaPPS40 solutions (10−5 mol L−1) in solvents with different polarities were studied (toluene, chloroform, tetrahydrofurane (THF), odichlorobenzene (o-DCB), and N-methyl pyrrolidone (NMP)) (Figure 9a). This figure shows that the higher energy π−π*

The photoluminescence (PL) spectrum for LaPPS40 in chloroform solution in Figure 10 shows a peak centered at λem

Figure 10. Electronic absorption and photoluminescence spectra of LAPPS40 in chloroform solution (10−5 mol L−1): (a) λexc= 445 nm and (b) λexc= 317 nm. Blue columns indicate the theoretically calculated oscillator strengths.

= 532 nm, independently of the excitation energy at λexc = 317 (which is resonant with the higher energy absorption) or 445 nm (with is resonant with the lower energy absorption). This indicates that when fluorene groups are preferentially excited there is a fast energy or charge transfer toward the lower lying excited state, which is responsible for the emission. This type of behavior has been reported for several types of copolymers containing donor and acceptor groups.10,39,40 In addition, the emission band is a mirror image of the absorption, with a Stokes shift of 3674 cm−1. This shift is significantly smaller than that observed for other conjugated polymers and copolymers, showing that the benzothiadiazole groups in the S1 electronic excited state do not undergo additional energy migration, energy transfer processes, or conformational relaxation.39,41 Thus, since no fluorene-related emission band is observed, the electron-deficient benzothiadiazole unit dominates the radiative process.42 The solvatochromic properties of LaPPS40 were further explored using steady-state photoluminescence in solutions with the same solvents used for the absorption. Figure 9b shows these emission spectra (10−4 mol L−1) (λexc = 445 nm). A stronger red shift was observed, according to increasing solvent polarity. Table 4 displays the results in numerical form. The correlation between the solvatochromic behavior and the solvent properties is in general very complex, but it can be described in a simplified form by the Lippert−Mataga equation, eq 4, under the approximation that the solvent forms a continuous dielectric field around the solute molecules26

Figure 9. Normalized electronic absorption (a) and photoluminescence spectra (λexc = 445 nm) (b) of LAPPS40 in different solvents (10−5 mol L−1).

transition is practically independent of solvent polarity, but the lower energy band shifts toward the red by 10 nm from toluene to NMP solutions (data in Table 3). The absence of spectral dependence of the higher energy band with the solvent polarity may be explained by the small changes of the dipole moments from 3.95 to 3.37 D (Figure 5) without significant change of its spatial orientation. Therefore, because there is only a small change of the electron density over the polymer chain we should not anticipate a major reorientation of the solvation layer around the macromolecule upon electronic excitation. On the contrary, for the S0 → S1 electronic absorption there is a spectral red shift in polar compared to nonpolar solvents, confirming the ICT character of this transition. Since the change in magnitude of the dipole moment is also small, this character must arise from the changes of its spatial orientation. Due to the change in charge distribution, the solvation layer must be reordered around the polymer chain after the electronic transition and this relaxation of the solvent leads to a larger red shift of the photoluminescence spectra.

Δν̅ = νabs ̅ − ν̅PL =

2⎤ ⎡ 2⎛ ε−1 n2 − 1 ⎞⎢ (μE − μG) ⎥ ⎜⎜ ⎟⎟ − 2 + cte ⎥⎦ hc ⎝ 2ε + 1 2n + 1 ⎠⎢⎣ a3

=

2⎤ ⎡ 2 ⎢ (μE − μG) ⎥ Δf + cte ⎥⎦ hc ⎢⎣ a3

(4)

where ν̅abs and νF̅ are the absorption emission energies in cm−1, respectively, h is the Planck constant, c is the light speed, ε for the dielectric constant, n for the refractive index of the solvent, μE and μG for the dipole moments of the electronic excited and electronic ground states, respectively, a for the radius of the solvent cavity where the solute molecule is located and νa̅ bs − ν̅PL for the Stokes’s shift. Δf is known as the solvent 3687

dx.doi.org/10.1021/jp211042v | J. Phys. Chem. A 2012, 116, 3681−3690

The Journal of Physical Chemistry A

Article

Table 4. Spectral Data for LaPPS40 in Solutions, Concentration 10−5 mol L−1, and Some Solvent Properties solvent properties toluene CHCl3 THF o-DCB NMP

Stokes shift

μa 43 (D)

n43

εb43

λabs(A1) (nm)

λabs(A2(ICT)) (nm)

λPL (nm)c

0.37 1.04 1.75 2.50 4.09

1.496 1.445 1.407 1.551 1.47

2.385 4.807 7.52 9.93 32.2

322 320 322 322 322

453 447 452 461 457

530 532 535 540 552

ν̅abs(A1) − ν̅PL (cm−1) 12 188 12 453 12 365 12 538 12 940

a μ is the dipole moment and n the refractive index. bε is the dielectric constant and ICT the internal charge transfer complex. cλexc = 445 nm. THF, tetrahydrofuran; o-DCB, o-dichlorobenzene; NMP, N-methyl pirrolidone. λabs (A1) = wavelength absorption of the first peak, λabs (A2) = wavelength absorption of the second peak.

polarizabity. Although this is a rough approximation for polymer molecules which are not of spherical shape in solution, Figure 11 shows that the Lippert−Mataga plot is approximately

Table 5. Emission Decay Components for LaPPS40 in Solutiona solvent

τ1 (ns)

B1 (%)

τ2 (ns)

B2 (%)

χ2

toluene 375 nm toluene 470 nm chloroform 375 nm chloroform 470 nm THF 375 nm THF 470 nm o-DCB 375 nm o-DCB 470 nm NMP 375 nm NMP 470 nm

0.755 ± 0.007

77

1.498 ± 0.031

23

1.001

0.759 ± 0.007

78

1.488 ± 0.03

22

1.000

0.964 ± 0.008

67

2.099 ± 0.02

33

1.000

0.904 ± 0.008

64

1.990 ± 0.02

36

1.000

± ± ± ± ± ±

73 71 80 76 64 53

1.982 1.868 1.795 1.639 2.413 2.129

± ± ± ± ± ±

27 29 20 24 36 47

1.000 1.000 1.000 1.000 1.000 1.000

0.915 0.880 0.858 0.811 1.084 0.942

0.007 0.007 0.007 0.007 0.009 0.009

0.03 0.02 0.04 0.03 0.02 0.02

Concentration 10−5 mol L−1; λexc = 375 nm, and λexc = 470 nm. λem = 550 nm.

a

obtained. It has been described that some conjugated polymers having ICT states may depict a clear trend of the emission peak position and lifetime with the solvent properties: solvents that bring about strong red shifts also induce a large increase in the lifetime.26,39 This is not the case of the LaPPS40 copolymer, probably because there is a very small change of the dipole moment when the charge transfer occurs from the fluorene to the benzothidiazole groups as shown by the quantum mechanical calculations. Biexponential decays as observed here may arise from several different possibilities. Since only dilute solutions were used, interchain interactions, interchain aggregation, and interchain excimer or exciplex formation were not considered plausible processes. However, intralumophore interaction brought about by chain folding cannot be disregarded, in principle. Indeed, the calculated geometry showed that the copolymer chain is not planar, and therefore, for longer chains, the possibility of close proximity between lumophores belonging to the same chain may be greater. Because of the sensitivity of the chain conformation to the polymer−solvent interactions, solvents with different dielectric constants should have a selective influence on the chain conformation and thus on the intrachain lumophore interactions. If this were an important contribution to one component of the biexponential decay, the relative contribution of the faster and longer components should change from one solvent to another. The absence of a significant dependence of the relative contribution of both components with solvent polarity suggests that also intrachain lumophore interaction is not playing an important role.

Figure 11. Lippert−Mataga plot showing the dependence of the Stokes shift with the solvent polarizability Δf (see eq 4).

linear (correlation coefficient equal to 0.8166 when NMP is included and 0.9955 when it is excluded). The solvatochromic properties using steady-state emission also gave some insights about the changes of the charge density along the benzothiadiazole groups, which have been confirmed by the dependence of the emission with the solvent polarity, but no information about the dynamics of the solvation can be inferred. For LaPPS40 in chloroform solution, the steady-state emission was also independent of the excitation wavelength, indicating that the charge transfer is a very fast process. The dynamics of the PL process was initially studied by recording the emission decay with laser pulses at λexc = 375 (where fluorene groups were excited) and 470 nm (where only benzothiadiazole units are excited) and collecting the emission signals at λem = 550 nm (data in Table 5). Data analyses were performed using multiple exponential functions according to eq 1, and good fits were obtained with biexponential functions for all solvents, with a predominant faster component (higher B values). There is no straightforward correlation between the decay values and the solvent properties (dielectric constants, Lippert−Mataga polarizability or solvent viscosity), although both decays for toluene (lower viscosity and lower dielectric constant) are faster than those for NMP. The contribution of the faster decay is always greater than the lower, but also no systematic correlation with the solvent properties has been 3688

dx.doi.org/10.1021/jp211042v | J. Phys. Chem. A 2012, 116, 3681−3690

The Journal of Physical Chemistry A

Article

Figure 12. Photoluminescence decays (acquision time 60 s) and time-resolved spectra for LaPPS40 in the range between 0.800 and 2.000 ns in toluene solution using λexc = 375 and 470 nm in several emission wavelengths.

Another possible explanation for the biexponential decay is that it arises from the presence of excited state lumophores in relaxed and unrelaxed CT states.40 This possibility might be considered taking into account that for all solvents the faster component is obtained when using lower excitation energy (470 nm) that corresponds to the red edge of the benzothiadiazole absorption band. Emission from the relaxed state should be faster than from nonrelaxed states, and thus, it has been assumed that the faster emission arises from the relaxed Franck−Condon excited state of the benzothiadiazole directly excited using 470 nm. The dynamics of the PL processes was also monitored for LaPPS40 in toluene solution using the same two excitation wavelengths (375 and 470 nm) and following the decay in the emission range from 400 to 700 nm (Figure 12). These decays were accumulated during 60 s. The decay patterns (left) are practically independent of the emission wavelengths where the decay is recorded. This is a typical behavior for localized lumophore excitation without spectral evolution by energy migration, energy transfer, or charge transfer processes, conformational relaxation, or strong solvent relaxation around the lumophore.26,39,40 Moreover, time-resolved spectra are also time independent and similar to those observed under steadystate conditions. Putting all the data together, the kinetic model for LaPPS40 decay seems to be very simple: when the donor is excited, it undergoes a very fast ICT process creating the electronic excited acceptor. However, because of the time resolution used in the present work is limited to 90 ps, we cannot go deeper into this mechanism. What can be said is that since the donor emission is absent the ICT process must be faster than the donor decay (around 100 ps to 1 ns for polyfluorenes). This requires an intrachain process with the donor and acceptor in close proximity, since the ICT process is very fast (